Some Historical Changes in the Patterns of Watershed

advertisement
American Fisheries Society Symposium 51:75–112, 2006
© 2006 by the American Fisheries Society
Some Historical Changes in the Patterns of
Population and Land Use in the Hudson River
Watershed
DENNIS P. SWANEY
Cornell University, Department of Ecology and Evolutionary Biology
Ithaca, New York 14853 USA
dps1@cornell.edu
KARIN E. LIMBURG AND KAREN STAINBROOK
SUNY College of Environmental Science and Forestry
1 Forestry Way, Syracuse, New York 13210 USA
klimburg@esf.edu
k_m_stainbrook@yahoo.com
Abstract.—Using a combination of data sources and historic or contemporary
accounts, we describe and document changes in the Hudson River watershed’s
population size, agricultural and forested land uses, and the construction of dams,
largely since the time of European colonization. Population within the watershed
has grown from 230,000 at the time of the first census in 1790 to around 5 million
today (not including parts of those boroughs of New York City outside the watershed,
such as Queens). The watershed was almost entirely forested in 1609, with minor
amounts of Indian agriculture. By 1880, approximately 68% of the watershed was
farmland, but as soil productivity declined and industry created other jobs, much
cleared land gradually reverted to secondary forest. Most land not in agriculture
was forested and exploited first for lumber and tanbark and, later on, pulpwood
for paper. The tanning industry existed from the 1700s, but reached its height in
the mid-1800s, collapsing from a combination of resource (hemlock) exhaustion
and market forces. Finally, available records list nearly 800 dams, ranging from
0.6 m to 213 m (Ashokan Reservoir) in height and with maximum storage of 1.07
km3 (Sacandaga Reservoir), that were constructed from the early 18th century
until 1993. The environmental legacies of these changes include effects on hydrology,
soils, vegetation, biogeochemical cycling, sediment loading, and ecological
relationships
Introduction
the countryside, eventually ponders the local history of the place and the effect of
earlier inhabitants on the local environment.
It is commonplace to discover the ruins of
homesteads and stone fences while walking
Anyone who has done environmental research in northeastern U.S. watersheds or
for that matter, has ever walked through
75
76
SWANEY ET AL.
in local forests and to observe the remnants of millworks and their impoundments in the waters of local streams.
Walking along the brackish river’s edge,
one finds oyster shells from beds dating
back before European settlement in
reaches where they are currently absent.
What are the connections between the
landscape, its flows of water, sediment
and nutrients, and the creatures that live
in its forests and waters? Such questions
prodded us to learn a bit more about
changes in the land use and land cover
of the Hudson-Mohawk watershed, and
to begin to assemble some data sets to
characterize these changes more fully. We
present some of this work below, with
the caveat that we are not historians and
so our efforts are subject to criticism by
those more expert in reconstructing the
historical record.
Our primary aim was to document
changes in regional land use noted in
census data at the county level and to
express these changes in terms of the corresponding change in the watershed and
subwatershed (hydrologic unit) land uses.
Thus, any discussion of changes in land
cover associated with aboriginal populations in the pre-European settlement period is largely beyond the scope of this
paper, and we refer the interested reader
to some previous excellent work done in
the northeastern United States (Cronon
1983; Foster 1999). Some work on the
historical reconstruction of pollutant loads
to the Hudson-Raritan has already been
done (cf. Ayres and Rod 1986; Rod 1986;
Rod et al. 1989). We focus primarily on
alterations caused by agriculture, forest
use, and dam construction, leaving such
topics as industrialization and urbanization for others.
Methods
County level census statistics were obtained
from the Bureau of the Census for both
population (U.S. Census Bureau, various
years) and agricultural land use (U.S. Department of Commerce, various years) in
New York and adjoining states with areas
in the Hudson-Mohawk watershed (i.e., New
Jersey, Connecticut, Massachusetts and
Vermont). U.S. Geological Survey (USGS)
8-digit Hydrologic Catalogue Units (HUC)
coverages of subwatersheds of the HudsonMohawk drainage were overlaid on county
boundaries obtained from the Cornell University Geospatial Information Repository
(CUGIR) using ARCView 3.2 (ESRI 1999)
to obtain the proportion of areas of each
county falling within each HUC and for the
entire Hudson-Mohawk watershed (Table
A.1). These proportions were then used as
multipliers on county population and land
use areas to determine area-weighted estimates for each HUC. A potential difficulty
with this approach is historical change in
county boundaries. While the county boundaries of the region have been largely stable
for the last 100 years or so, the period from
the first national census (1790) to 1910 saw
numerous changes in county boundaries,
including the formation of new counties at
the expense of old ones. We used the Historical U.S. County Boundary (HUSCO)
data set to evaluate areas of counties for
the years of the decennial census in this
period (Earle et al. 1999). The period between the earliest settlements (1600s) to the
first national census saw even more radical
changes in political boundaries, but change
in this period is largely beyond the scope of
quantitative analysis of this work. For census years in which county boundaries are
very different from current boundaries, areaweightings had to be calculated separately,
as shown in Table A.2 of the Appendix.
HISTORICAL CHANGES IN POPULATION AND LAND USE
To assess changes in population and population density, we used county level data
from the decennial U.S. census records and
the New York State Data Center (New York
State Data Center 2000; U.S. Census Bureau, various years), from which total population counts of each county were obtained.
Similarly, the Censuses of Agriculture were
used to determine changes in acreage of land
in farms over time. Figures on county and
subwatershed population and land in farms
are given in Tables A.3–A.6 of the Appendix.
Forested land area by county is more difficult to obtain and was determined from
several different sources, depending upon
the period of interest. These included Hough
(1878) for the year 1875, Recknagel (1923)
for the year 1920, Stout (1958) for the year
1955, USDA Forest Resource Bulletins NE20, NE-34, NE-44, NE-46, and NE-48 for
the period 1968–1972, and Forest Resource
Inventory Reports NE-112, NE-132, NE145, NE-147, and NE-148 for the late 1980s
through mid-1990s. These latter inventories are updated approximately every 10
years, rotating through the states within a
jurisdictional region of the U.S. Forest Service. Forest land areas in counties and
subwatersheds are shown in Tables A.7A.8 of the Appendix.
The U.S. Army Corps of Engineers maintains the National Inventory of Dams (NID)
for all dams greater than 2 m in height
with a storage greater than 61,700 m3, or
greater than 8 m in height with a storage
greater than 18,500 m3, or of any size likely
to pose a significant threat downstream
(Graf 1999). The U.S. Environmental Protection Agency incorporates this information into its Better Assessment Science Integrating Point and Nonpoint Sources (BASINS) software for watershed analysis
(USEPA 2001). For these analyses, we extracted the dam data for the Hudson-
77
Mohawk basin. Dam data organized by
subwatershed and size are shown in Table
A.9. The NID includes the dates of dam
construction.
It is worth noting that the watershed is not
a natural unit for the study of populations,
and in the case of the Hudson, the estimate
of “watershed population” is heavily
weighted by the counties nearest New York
City. As a result, for these counties, even
small errors resulting from the assumption
of uniform spatial distribution of population, or from the proportion of county areas
falling within the watershed boundaries, may
have significant effects on the total population estimates. However, our aim is not to
further the analysis of New York demographics, but rather to assess the spatiotemporal
variation of major environmental drivers,
and as of this writing, we believe the use of
county level data for this purpose is the
state of the art.
Results and Discussion
Pre-European Settlement and Colonial
Periods
“There was not an unbroken forest here when
the first settlers came; as the fires of the
Indians, in their pursuit of game, had destroyed the timber on the dry lands, except a
few specimens of oak, white wood and wild
cherry, some of which attained great size.
On the plains were scattered small oaks which
had sprung up after the fires, and by the
creeks and in wet lands there were large buttonwood and black ash trees, while all the
streams were overhung with a mass of alders
and willows. The mountains, it has been said,
were covered with a less dense growth of wood
than at present. It is evident that in the
valleys, the white wood or tulip tree, and the
wild cherry have given place to other trees,
78
SWANEY ET AL.
Figure 1. Distribution of primeval forest cover in the region (redrawn from Hamilton et al. 1980).
Alleghenian hardwoods include primarily beech, sugar maple, hemlock, white pine, and basswood;
Adirondack hardwoods are primarily beech, sugar maple, yellow birch, hemlock, and white pine.
as the elm; and that on the mountains, the
chestnut has greatly increased. The mountains, being burned over also by the Indians, were so bare, that the wild deer were
plainly seen from the valleys below.”
- Early History of Amenia. Newton Reed,
1875.
Aboriginal peoples have been present in
the area for at least the last 7,000 years, at
first as hunters and gatherers, and later
creating semi-permanent settlements and
practicing agriculture (Hamilton et al. 1980;
Cronon 1983). As the above quote suggests, lands of the watershed were largely
forested (Figure 1) with the exception of
areas cleared by the inhabitants, often by
burning, in the process of hunting and
planting crops (Day 1953).
The earliest European settlement in the
Hudson region was New Netherland, established by the Dutch West India Company,
a follow-up to the Dutch East India Company that had sponsored Henry Hudson’s
1609 searches for a northwest passage to
Asia (Shorto 2004). New Amsterdam, its
main population center, was established first
as a fort at the southern tip of the island of
the Manhattes (Lenape) Indians in 1623;
Fort Orange (Albany) was established that
same year. Peter Minuit, Director of the
New Netherland colony from 1626–1633,
purchased Manhattan Island to gain greater
control over it. Early settlement patterns
followed the main waterways (Figure 2), as
these were the main means of transporting
goods through the region. Settlement up
the Hudson Valley was slow and interspersed
HISTORICAL CHANGES IN POPULATION AND LAND USE
79
Figure 2. Patterns of settlement in the watershed (redrawn from Fox, 1900).
with conflicts with both English and Indians. Eventually, the Dutch yielded their
possessions to the English in 1664, and New
Amsterdam was renamed for the Duke of
York.
Just as with the English settlers in New
England, the Dutch were impressed by the
abundance of wildlife, the plenitude of forests, and the beauty of the landscape.
Montanus (1671) described in some (often
fabulous, Figure 3) detail wolves, lions (panthers), black bear, deer, elk, ”muskcat“
(probably muskrat), beaver, rattlesnakes,
and many kinds of waterfowl and birds of
prey. Among the birds of the forest, he noted
that ”pigeons fly in such flocks that the
Indians designedly remove to their breeding places, where the young birds, pushed
by hundreds from their nests, serve for food
during a long month for the whole family.“
Turkey was in great abundance, and the
streams were filled with many fish species.
Boyle (1969) lists many similar accounts.
From the start, tree clearing was necessary
to ”improve the land“ and establish agriculture. To this end, the West India Company constructed three sawmills in its new
territory (Anonymous 1647). With machinery imported from Holland, these were constructed to operate with either wind or water power (Fox 1900). One was located on
Governor’s Island and another on Sawmill
Creek, draining into the East River (Fox
1900). Other sawmills were constructed for
purposes of clearing in the 1600s in Albany, Westchester (on the Sawmill River),
Kinderhook, Troy, Castleton, and Catskill
(Fox 1900). Trees were girdled and cut with
axes, and wood was used for buildings, fence
rails, and posts. A great deal of wood in
the early settlements was simply burned to
clear the land.
SWANEY ET AL.
80
Figure 3. Wild beasts of New Netherland, depicted by Montanus (1671). Note the mixture of real and
fabulous Hudson Valley species.
Population
Population growth was slow during the period of Dutch settlement (1600s) and English colonization up through most of the
18th century, as large land grants and patents had been given to a few families such
as the van Cortlandts, the Livingstons, the
Beekmans, and the van Rensselaers, who
maintained near-feudal conditions in the
early part of that century (O’Callaghan 1850;
Ellis et al. 1973). After the American Revolution, watershed population grew steadily.
Population growth was exponential from
1790, when the first national census data
available indicate about 230,000 people in
the watershed, until around 1920, when the
population reached about 3.4 million (Figure 4a). The engine of population change
for the watershed has always been New York
City. While the total watershed population
is dominated by that of the metropolitan
area (here, seen as the lower Hudson HUC,
Figure 4b), population farther north in the
watershed has always followed the transportation corridors.
Construction began on the Erie Canal in
1817; it was opened in 1825, and development of feeder canals and enlargements continued through the 19th century, facilitating trade and settlement in rural areas. By
the 1840s, railroads had begun to change
population distributions in the counties
along the main trunklines paralleling the
Hudson and expanded the effective metropolitan area of New York City. Most of the
New York State population today falls in
81
HISTORICAL CHANGES IN POPULATION AND LAND USE
Hudson Watershed Population over time
5
(Millions)
Population
4
3
2
1
0
1790
1840
1890
1940
1990
Year
Figure 4a. Population change in the Hudson watershed, 1790-2000.
Figure 4b. Breakdown of change by subwatershed. Data source: U.S. Census, county level data.
82
SWANEY ET AL.
Figure 5. Changes in spatial distribution of population density over time at the county level.
a band along the Hudson/Mohawk rivers,
and extending westward along the transportation routes, which lie along the old
canal beds to Buffalo (Ellis et al. 1967).
Between 1920 and 1930, a period which
saw the beginning of the Great Depression,
the watershed population appears to have
declined (Figure 4a, b). While it is worth
noting that in the fall of 1918 the country
experienced the great Spanish influenza pandemic, in which 675,000 individuals lost
their lives in the United States, and over
50,000 in New York State (Brainerd and
Siegler 2002; Eichel 1923), the pandemic
was not a major driver of population trends.
Rather, the decline coincides with barriers
to the influx of foreign immigrants raised
by the Immigration Restriction Act of 1921
and the Immigration Act of 1924 (Ellis et
al. 1967). Population shifts were also due
to changes in the economy: as regional
agriculture declined, rural population in the
upper watershed moved to cities and suburbs as people looked for better economic
opportunities. In the lower Hudson, population has remained steady with decadalscale variation over the last hundred years,
while from around the 1930 census, overall
watershed population has grown steadily but
at a lower rate than in the previous century,
to around 5 million today.
Agriculture
Agriculture has been important in the watershed since well before European settlement. Aboriginal inhabitants planted corn,
beans, and squash and tapped maple forests for sugar (Hedrick 1933; Ellis et al.
1967). The Dutch settlers of the precolonial
period planted wheat for export, and this
remained a major cash crop in the region
83
HISTORICAL CHANGES IN POPULATION AND LAND USE
Land in farms in Hudson Watershed over time
100
30000
60
20000
40
Watershed area
10000
Watershed %
NY state %
0
1850
% Land in farms
2
Land in farms (km )
80
20
1900
1950
0
2000
Year
Figure 6. Agricultural land use in the watershed peaked in the 1870s-1880s. Data source: U.S. Census
of Agriculture, county level data.
through the first decades of the 1800s when
it became unprofitable because of the effects of the Hessian fly, wheat rust, and
other insect and disease pests (Hedrick
1933). By the 1850s, dairy farming had surpassed wheat in economic importance in
the state. Corn for livestock feed and other
small grains replaced wheat as it became
less profitable (Ellis et al. 1967). Land in
farms grew steadily in the early part of the
19th century and apparently peaked in the
watershed around 1880 at about 68% of
the total area (Figure 6), about 11% less
than the corresponding peak of the state
overall. During the peak years of agriculture in the watershed, land in farms exceeded 90% in several counties, though it
is important to note that farmland included
”unimproved land“ (woodlots, etc), and trees
harvested for lumber and other uses were a
significant source of farm income, especially
early in the 19th century. The trend thereafter until the mid-20th century is negative—
a steady decline until the late 1960s and a
less steep decline from then until the present
day value of around 15% of the watershed
area. A slight upward trend in the late 1940s
reflects agricultural demands of the war
years. This pattern has reflected that of New
York State to a remarkable degree, though
the proportion of watershed land in farms
has been somewhat lower than the corresponding figure for the state as a whole for
at least the last 150 years (Figure 6).
Agricultural land uses have never been uniformly distributed over the watershed (Figure 7). Orchards have thrived along the
Hudson from the days of earliest settlement.
The New York dairy industry began in
Dutchess, Herkimer, Oneida, and Orange
counties in the early decades of the 1800s
and spread to other counties of the watershed within the next 30 years (Hedrick 1933).
Beginning in the 19th century, some areas
in the watershed became known for specialty crops (flax in the Hoosic Valley, hops
in Otsego County). Rye was grown in the
84
SWANEY ET AL.
Figure 7. Spatial distribution of agriculture as percent of land use in the watershed, 1875-1997. Data
source: U.S. Census of Agriculture, county level data.
Hudson Valley to meet the demand of distilleries (Ellis et al. 1967). At the peak of
agricultural intensity, over 90% of the land
area of counties along the Hudson River
corridor was allocated to farms (Figure 7).
The northern counties of the watershed remained relatively unforested at that time,
as did the southernmost tip of the watershed in New Jersey, though for very different reasons. Logging operations and associated fires deforested large areas of the
Adirondacks, Catskills, and Hudson River
Highlands (Fox 1900; Moon 1909; see next
section). The southernmost tip of the watershed was experiencing explosive population growth as the New York City metropolitan area expanded outward in response
to migration from other regions of the United
States as well as immigration from abroad
(Ellis et al. 1967).
regional urban and suburban populations
have increased at the expense of rural populations throughout the 20th century (Ellis
et al. 1967). Today, while the counties of
the region all have substantially less land
in farms, the distribution reflects that of
the height of agricultural activity in the region: counties with the greatest proportion
of agricultural land tend to follow the main
river corridors. Land in the north is largely
reserved as parkland, and the watersheds of
the New York City water supply system in
the Catskills and the counties immediately
north of New York City are controlled to
some extent by regulations of the New York
City Department of Environmental Protection. The suburbs of the New York City
metropolitan area have grown to cover most
of northern New Jersey and the southern
counties of the watershed.
As in most other parts of the United States,
Finally, while a detailed discussion of the
HISTORICAL CHANGES IN POPULATION AND LAND USE
effects of agricultural technology is beyond
the scope of this study, a major aspect of
20th century agriculture was the enormous
growth in pesticide and fertilizer use, especially following World War II. In the HudsonRaritan basin, organochlorine pesticide loading grew from around 1,400 kg in 1945,
peaking in the early 1970s (12,400 kg in
1971), and then declining in the 1980s (Ayres
and Rod 1986). The total nitrogen load from
fertilizer in the counties of the Hudson/
Mohawk watershed also grew rapidly from
the 1940s (5,000 metric tons N) until the
mid 1980s (28,000 metric tons N), but has
declined, with fluctuations, since then
(Alexander and Smith 1990; Battaglin and
Goolsby 1994). It should also be clear that
conventional agriculture is not the only
source of fertilizers and pesticides; beginning in the 1950s, commercial and residential lawns have rivaled row crop agriculture
in the intensity of pesticide and fertilizer
use, though quantitative load estimates are
more difficult to assess.
Forests
The history of New York, including the
Hudson River watershed, is intimately tied
to its forests. After the last glaciation, succession resulted in an oak and chestnut dominated forest in the Hudson Valley, northern
hardwoods (composed primarily of beech,
maple, yellow birch, hemlock, and white
pine) in the Catskills and lower Adirondack
regions, and a spruce, fir, and paper birch
complex dominated the higher elevations of
the Adirondacks and Catskills (Hamilton
et al. 1980). Indians of the region manipulated the forests to enhance productivity of
resources, girdling trees, and burning areas
to open up the woods for wildlife and creating habitat for fruit-bearing plants (Day
1953). When early European settlers arrived,
they encountered large areas in the forests
of southern New England and the lower
85
Hudson that were “remarkably open, almost parklike at times” (Cronon 1983; p.
25). The northern forests were denser.
As suggested above, the early settlement
and colonial days were largely periods of
land clearing in which Dutch, and later
English, settlers cleared land for farms and
established sawmills along local streams.
Revenues from wood exceeded those from
crops. Early lumbering, beyond the needs
of the colonials, sought out first the ancient
white pines that were used by both the
Dutch and English royal navies as ships’
masts and timbers. Fox (1900) noted that
the tallest pines were about 255 ft (89 m)
tall and over 2 m wide and that Pine Street
in New York City was named for the ”many
magnificent pines“ on the farm of January
Jansen Damen. In addition to white pine,
oak and hickory were exported, first to
Amsterdam and later to England, which
had suffered wood shortages since the 1500s
(Cronon 1983).
After the American Revolution, and consequent routing of the Iroquois, the colonists
expanded northward and westward. As demand for wood increased, sawmills were
erected wherever sufficient supplies of wood
and water occurred (Hamilton et al. 1980).
Logs cut close enough to the sawmills were
dragged in by draft animals, but as trees
became depleted, the mills could be moved
to other sites with more trees. In this manner, forests were selectively cut for decades,
with mills successively moved to more remote places.
Tanning was a parallel, major forest-based
industry that arose early on in the Hudson
Valley and had a significant impact on forested ecosystems. Colonial tanning operations began in New England, using oak,
chestnut, and hemlock bark as the source
of tannins. As the supplies dwindled there,
the industry pushed westward into the
86
SWANEY ET AL.
Hudson Valley, eventually heading up into
Michigan, Wisconsin, and ultimately the
West Coast. By 1810, there were 867 tanneries in New York State (McMartin 1992),
many of them in the Hudson Valley and
most of them small scale. However, the large
financial interests of “The Swamp,” a group
of New York City leather businessmen named
for that section of the city where tanning
was practiced, selected and backed the development of large-scale operations, first in
the Catskills and later in the Adirondacks.
Colonel William Edwards, a tanner and
grandson of the famed Boston minister
Jonathan Edwards, not only received financial backing from The Swamp but also lobbied to have the laws changed to allow tannery corporations to operate in Green and
Delaware counties (but not Ulster and
Sullivan, where anti-tanning sentiments were
higher; Evers 1972). His New York Tannery
opened in 1816 in Edwardsville (now
Hunter). An 1820 painting depicts the tannery by the Schohariekill, showing the onceforested land laid to waste all around (Evers
1972). Zadock Pratt, for whom Prattsville
is named, developed the prototype tannery
that became common in the Adirondacks
(McMartin 1992).
Hemlock came to be the preferred resource,
as it grew in great stands on north-facing
slopes in the Catskills and on lower slopes
in the Adirondacks. Tanning used only the
hemlock bark and not the rest of the tree,
which was of little value for timber. Thus,
it was a wasteful industry, with trees felled
and only the bark peeled for use. The stems
were left to decay, and were described as
looking like the bleached bones of giants.
Tanning also required a plentiful supply of
water to run the operations and into which
to sluice the wastes, which were considerable. The putrid smells and severe organic
pollution of tanneries ruined many rivers
and streams while they operated. As the
Catskills came to be depleted of hemlock,
operations moved into the Adirondacks. A
general trend from small-scale to increasingly larger operations took place from the
1830s until the end of the tanning era in
New York in the 1890s (McMartin 1992).
Finally, as the hemlocks became increasingly scarcer and newer technology came
online, tanneries moved closer to the sources
of leather and the industry collapsed in New
York.
While the tanning industry declined, logging for timber and pulp continued to grow
throughout the 19th century. Rafts were used
to float sawed timber down to markets to
the north, west, and southeast (Fox 1900).
Because of the tides, rafts could not be
floated on the Hudson below Albany, and
so the wood was loaded onto ships that
sailed down to New York City. Log driving
was commenced in 1813 to send timber
down to the mills in Glens Falls. This innovation was quickly repeated and became
the main means of getting logs to the mills.
Throughout the early to mid-1800s, rivers
became declared public transportation highways through acts of the legislature. Lake
levels throughout the Adirondacks were
raised with “spill-dams” that were removed
in spring to augment the freshets and move
the logs downstream. The towns of Glens
Falls, Sandy Hill, and Fort Edward became
the center for the emerging wood products
industry. In 1849, the Hudson River Boom
Association set up “booms” or collecting
and sorting points (Figure 8). Logging companies would place unique brands on their
logs to identify them. Many accounts exist
documenting the bravery of the log drivers,
who would risk their lives breaking up logjams.
The advent of rail provided a new means of
moving timber out of the Northern Forest.
The first railroad into the Adirondacks was
opened in 1868 (Donaldson 1921). Although
railroads were built at first to move in equip-
HISTORICAL CHANGES IN POPULATION AND LAND USE
87
Figure 8. Photograph of the Glens Fall Boom holding logs from the Northern Forest. Source: Fox (1900).
ment, they were soon put to use in transporting out wood (Hamilton et al. 1980).
Rail carried much of the hardwood that was
too heavy to float down rivers and thus
facilitated the exploitation of relatively unused species (Fox 1900). Steam engines, with
hot coals and sparks flying, were also responsible for many of the forest fires that
plagued the region.
increasingly less selective. Timber production peaked in the 1870s, and pulpwood
production peaked in 1908 (Canham 1981).
Although quantitative data were unavailable prior to 1870, the time trend (Figure 9)
shows that 20th century wood production
was never more than half of the 19th century highs, again suggesting the severe
overexploitation of the forests in the 1800s.
In 1867, the sulfite process for reducing the
fiber from wood pulp (produced by grinding wood) was introduced, and wood fiber
became a new source of material in papermaking, which had previously been made
with rag and cotton fibers (Recknagel 1923).
As this industry began to take off, smaller
trees could be exploited. Lumbering became
The history of forest use in New York up to
the late 1800s was a picture of exploitation
and general lack of concern for the welfare
of the forested ecosystems. Forest destruction began to raise fears of large-scale environmental degradation. Early environmentalists noted the increased sediment loads
in the Hudson (Hamilton et al. 1980) and
SWANEY ET AL.
88
Lumber production in NYS, 1870-1978
10
Hardwoods
Softwoods
Tons of lumber
(Millions)
9
8
7
6
5
4
3
75
19
65
19
55
19
45
19
35
19
25
19
15
19
08
19
00
19
90
18
80
18
18
70
2
1
0
NY State wood products statistics (data: H.Canham, 1981)
Figure 9. Temporal trends in production of lumber and pulpwood in the Northern Forest. Data source:
Canham (1981).
streams, warmed from lack of shade, were
losing their trout and other coldwater species. At the same time, forest scientists in
Europe were becoming aware of the connections between deforestation in the Alps and
the lowering water level of Lake Geneva
(Hough 1878), and the nascent field of scientific forestry began prescribing new practices to reduce the environmental impacts
of the industry. The depredations of logging in New York inspired a movement to
preserve the forest, lest it be completely ruined. Spurred on by the eloquent lobbying
of Verplanck Colvin, the state surveyor who
mapped the Adirondacks, in 1885 the legislature enacted a law to establish permanent
forest preserves and develop a system of
forest fire protection. Although this “forever wild” law prevented the sale of lands,
it did not prevent tree harvesting. The preserve began with the acquisition of tax-delinquent lands, and from the 1890s onward,
also began a program of land purchasing
and reforestation. This activity was administered by the newly consolidated Forest,
Fish and Game Commission, which worked
to train a new generation of “scientific” foresters. The work of the Commission also
involved fire prevention and fire fighting.
Droughts, in combination with the practice
of leaving intact treetops after cutting the
logs, promoted fire hazards; sparks from
trains, lightning, from careless campers
caused massive fires in the early 1900s (Figure 10). The Commission’s reports motivated legislation for preventing fires, with
stiffer fines, modification to the rail rightsof-way, and fire towers as some of the innovations (Whipple 1909). Another preserve
in the Hudson River Highlands was created
in 1909, in response to extensive clear-cutting for fueling the brickyards, as well as
susceptibility to fires created by poor stand
conditions (Moon 1909).
HISTORICAL CHANGES IN POPULATION AND LAND USE
89
Figure 10. Causes of forest fires, from the 1908 Report of the Forest, Fish and Game Commission.
Severe droughts were associated with the high-fire years (symbols have been added to the original figure).
90
SWANEY ET AL.
Figure 11. Spatial distribution of forested land as percent of land use in the watershed, 1875-1990.
Data sources: Hough (1878), Recknagel (1923), Stout (1958), and U.S.D.A. Forest Resource Bulletins and
Forest Resource Inventory Reports.
Better management practices in the 20th century promoted regrowth of many of the forests. Forests were also managed for watershed protection, with recognition of this as
early as the first report of the Forest, Fish
and Game Commission. Programs of woodlot
management were begun and carried out by
farmers with help from state foresters
(Hamilton et al. 1980). Forest inventories,
carried out every decade by the U.S. Forest
Service, show a pattern of recovering forests
throughout the Hudson basin counties into
the 1990s (Figure 11). This pattern was also
driven in part by the abandonment of farmland, as agrarianism gave way to a fossilfuel supported economy, with resources increasingly drawn in from outside.
Dams in the Hudson Watershed
Dams were likely among the earliest modifications of streams within the Hudson watershed, as waterpower was needed for driving sawmills and other small industry. Dams
were constructed for a number of other purposes, including tanning operations, raising water levels to ensure adequate flow for
log drives (see Forests section) and for feeder
canals along the Erie and Champlain canal
systems, flood control, water supply reservoirs, and hydropower. Inevitably, these systems often serve multiple purposes and, in
age (e.g., the Eddyville Dam on the Rondout
Creek), sometimes outlive their main purpose.
HISTORICAL CHANGES IN POPULATION AND LAND USE
Figure 12. Accumulation of dams over time in the Hudson watershed. Top four maps show the
cumulative change in numbers of dams, by dates of completion, every 50 years from 1850 onwards.
Bottom map shows all dams, inclusive of those lacking dates. Data source: BASINS (USEPA 2001).
91
92
SWANEY ET AL.
The BASINS data set includes the dates of
completion of most dams in its database,
but 128 dams (16% of the database for the
Hudson drainage) lack dates. Some of these
may well be among the earlier dams. Furthermore, the database is not an exhaustive
list of all dams in the watershed (Graf 1999).
Nevertheless, the spatiotemporal trend of
dam construction (using records with dates)
is dramatic (Figure 12) and appears to capture the general pattern of the entire data
set. The period of 1900–1930 was an age of
rapid dam construction in the Hudson and
presaged the era of great dam construction
in the West. Many of the largest dams were
built in this period in the Hudson (Figure
13A), including the Kensico and Ashokan
Reservoir dams in 1916 with a combined
storage capacity of 0.8 km3 (629,060 acrefeet) and the Sacandaga Reservoir, completed in 1930 both for flood control and
hydroelectric generation, with a displacement of 1.07 km3 (866,000 acre-feet). Examining the rate of change of cumulative
storage capacity (Figure 13B), it appears
that most of the damming potential of the
basin was built out before 1950, with relatively minor additions since then.
Today there are some 797 dams within the
Hudson watershed proper, with another 357
extant in the surrounding basins of the
Hackensack-Palisades, Sandy Hook, Raritan,
and Bronx. Most of the dams range in height
between 2 and 10 m (Table A9), the tallest
being the Ashokan Dam at Olive Branch
(213 ft, 64.9 m). The “damming intensity”
(number of dams per subcatchment area) is
greatest in the Lower Hudson (0.069 km-2 or
0.18 mi-2) and lowest in the Sacandaga (0.011
km-2 or 0.03 mi-2).
We note that the database likely does not
include very small dams (ca. 0.5 m tall)
and impoundments and that these could
also be abundant and significant in the overall hydrology of the watershed. Even with-
out including these effects, when the cumulative impoundment is expressed in units
of average annual flow of the Hudson at the
Troy dam (Figure 13B), it appears that impoundments represent the equivalent about
4 months of flow. If we take this as an
average of the effect on streams throughout
the watershed, dams currently have the capacity to impound one-third of the annual
flow of water through the catchment and
consequently retard and dampen the response
of streamflow to precipitation events.
Hydrological, Biogeochemical, and
Ecological Implications of Changes
in the Hudson/Mohawk Watershed
Clearing the forests of the watershed for
agriculture undoubtedly had the most dramatic impact on the watershed as a whole.
Some simple modeling results have suggested that the fully forested primeval watershed would have seen erosive soil losses
around one-eighth of current rates and, during the peak days of agricultural activity of
the 1800s, more than double modern rates
(Swaney et al. 1996). The combined effect
of clearing, burning, and then tilling the
soils poured sediments into the streams,
removed the litter and organic matter, and
eventually wore out some of the soils. Some
simulations of agricultural landuse in the
northeast United States have suggested that
long-term depletions of soil N and C, due
to the disturbance of biomass removals, are
greater than losses associated with timber
harvesting (Ollinger et al. 2002). Early leaders of agriculture recognized that careless
agricultural practices depleted the soils of
nutrients and complained of the effects of
poor management on the condition of soils
and livestock of the region (e.g., Ellis et al.
1967; quoting G. W. Featherstonhaugh
1819), and the 19th century migration of farmers from the region to the western frontier
was largely a matter of literally moving to
93
HISTORICAL CHANGES IN POPULATION AND LAND USE
70
Dam height (m)
60
A
50
40
30
20
10
0
1750
1800
1850
1900
1950
2000
2050
Year of completion
B
0.3
3
Cumulative storage (km )
3.5
0.35
3.0
0.25
2.5
0.2
2.0
0.15
1.5
0.1
1.0
0.5
0.05
0.0
1750
0
1800
1850
1900
1950
2000
Cumulative storage - annual
flow equivalents (yr)
4.0
2050
Year of completion
Figure 13. Time series of (B) the heights and (B) the cumulative storage capacity of dams in the HudsonRaritan drainage. Data source: BASINS (USEPA 2001).
greener pastures. While tile drainage of agricultural fields began in the United States
in Seneca County, New York, to the west of
the Hudson watershed, as early as the 1830s,
farmers in several counties within the wa-
tershed had adopted this technology by the
1860s (Weaver 1964). Today, the benefits of
tile drainage to crop yield in wet, heavy
soils are recognized, as are the consequent
increases in water discharge and agricul-
94
SWANEY ET AL.
tural chemical transport from fields (McIsaac
and Hu, 2004).
Changes in the forests had their own set of
impacts. The early high-grading practices
for white pine at first, and then other species, ecologically represented selective species removal, which would have implications both for biogeochemistry (e.g., Hobbie
1992; Chapin et al. 1997; Lovett et al. 2002)
and forest community dynamics (e.g.,
Canham and Pacala 1995; Pacala and
Deutschman 1995; Chapin et al. 1997).
Similarly, the practice of cutting hemlock
trees, removing the bark, and leaving the
stems to decay might have had effects similar to those shown at Hubbard Brook Experimental Forest watershed 2, in which
stream nitrate concentration increased significantly in response to trees being cut
and left in place (Likens and Bormann
1995). The effects of forest harvest practices, fires, and land use change on forest
biogeochemistry are active areas of ecological
research (e.g., Likens and Bormann 1995;
Aber and Driscoll 1997; Waring and Running 1998; Goodale et al. 2000; Goodale
and Aber 2001; Latty et al. 2004). Generally, biogeochemical functions become
“leakier” in the years immediately after
harvesting and especially after fires; other
effects can be detected more than a century after the fact (Goodale et al. 2000;
Goodale and Aber 2001; Latty et al. 2004).
Many other factors, such as geology, stand
age, and forestry practices may affect the
particular direction of change, depending
on circumstances (e.g., Yanai et al. 2000;
Lovett et al. 2002). Changes in species composition could also have implications for
the quality of allochthonous materials delivered to lake, stream, and ultimately estuarine food webs (Pace et al. 2004).
Removing trees also affects hydrology, by
altering evapotranspiration among other
things. Although transpiration is lowered,
openings in the forest can warm the soil
and increase evaporation. The foresters of
the late 19th and early 20th centuries were
well aware of the connections between forest cutting and hydrology (Hough 1878;
Newell 1900), and Verplanck Colvin used
these arguments as part of the justification
to establish the Adirondack Preserve (Fox
1900; Hamilton et al. 1980). The early fisheries scientists were also cognizant of the
connection between forest cover and moderation of streamflow and temperature.
Charles Stevenson (1899), writing about the
status of the anadromous American shad
Alosa sapidissima, drew the connection between deforestation, agriculture, siltation,
increased temperatures, and the decline of
this marine-dwelling animal. He wrote:
“During heavy rains the plowed soil upon
the hillsides is easily washed into
gullies…[and] quickly conveyed to the rivers, filling them beyond their capacity and
bringing into them masses of earth and other
debris, thus covering the spawning grounds
(p. 113).” The great Adirondack droughts
and forest fires of the early 20th century
may even have affected the water temperature of the tidal Hudson River as far downstream as Poughkeepsie, causing a warming of a few degrees at that time (Abood et
al. 2006, this volume).
While we have not considered explicitly the
changes in the urban landscape and road
network within the watershed, it is easy to
demonstrate that population density can
serve as a proxy variable for road density,
as they vary in space (e.g., county to county)
and over time. Consequently, it is apparent
that the changes in population density within
the watershed indicate corresponding
changes in road and building density and,
therefore, impervious surface density. The
resulting implications for runoff and water
quality are obvious: faster responses to precipitation and snowmelt events, higher road
salt loads in winter and spring, elevated
HISTORICAL CHANGES IN POPULATION AND LAND USE
water temperature from warm pavement in
the summer, and increased dissolved and
particulate substances related to automobile traffic. On the other hand, the high
organic nutrient loads that were ubiquitous
in the roads and streets of the “horse and
buggy days” have essentially vanished.
Modern transportation networks affect more
than just the hydrology and chemistry of
the local environment, serving as both barriers to wildlife movement and conduits for
invasive species, increasing wildlife mortality, and so forth. Forman and Alexander
(1998) suggest that a road density of 0.6
km/km2 is the maximum for a “naturally
functioning landscape.” According to New
York Department of Transportation (DOT)
figures, of all the counties falling all or partly
within the watershed, only two met this
criterion in 2001: Essex and Hamilton, in
the Adirondacks (New York State Department of Transportation 2002).
The environmental impacts of dams receive
much attention today (McCully 1996). The
alteration of hydrological regimes is an obvious effect, and large dams worldwide have
been documented to slow the release of waters draining off the continents, with concomitant increased evaporation (Vörösmarty
et al. 1997). Even small dams, when taken
in the aggregate, may have measurable hydrologic effects. Smith et al. (2002) note
that small impoundments in the contiguous United States, most of them in the East,
account for 20% of the standing water area
and affect hydrology, sedimentation, and
biogeochemistry. Dams cause increased
sedimentation upstream in reservoirs and
increased erosion downstream. Dean and
Gorham (1998) estimated that reservoirs
worldwide have sedimentation rates 4.7-fold
greater than lakes. It is also recognized that
many biogeochemical processes (e.g., denitrification and settling of particulates) are
sensitive to the residence time of waters flow-
95
ing through the landscape; longer residence
times result in more settling and nutrient
“processing.” A model-based analysis of the
effect of reservoirs on denitrification by
Seitzinger et al. (2002) found very little impact at the catchment scale, in part because dams are often sited upstream of nitrogen sources in watersheds. On the other
hand, Humborg et al. (1997) found dams
not only to affect river biogeochemistry (e.g.,
to alter dissolved nutrient ratios), but to
propagate effects on ecosystem structure and
productivity in recipient marine waters.
Finally, dams can have devastating ecological effects, which have been best documented
on fisheries. Dams have contributed to the
downfall of Pacific salmon through the loss
of spawning habitat (Nehlsen et al. 1991),
and Limburg et al. (2003) estimated that
over 4,000 km of American shad spawning
habitat have been lost due to damming
North American East Coast rivers. A number of studies have documented the potentially important links of the transport of
marine-derived nutrients to freshwater ecosystems (e.g., Richey et al. 1975; Durbin et
al. 1979; Bilby et al. 1996; Garman and
Macko 1998; MacAvoy et al. 2000), providing both direct and indirect subsidies to
stream and riparian zone ecosystems; these
links are generally truncated by dams.
Pringle (1997) and Freeman et al. (2003)
document the disruption by dams of many
other ecological interactions, and even shortterm evolution has been documented by the
splitting of populations by dams (Morita
and Suzuki 1999). At this point, we can
only guess at the environmental and ecological changes effected by dams in the
Hudson-Mohawk watershed, but undoubtedly, given the degree and magnitude of
damming, the effects have been considerable.
Table 1 summarizes the environmental impact of various human activities in the wa-
SWANEY ET AL.
96
Table 1. Changes in watershed characteristics and associated effects.
Pre-European settlement
Aboriginal agriculture
Localized fires and associated change in
biomass, habitat, and nutrient
remobilization
Pre-Colonial and Colonial settlement
Clearing of land for agriculture
Removal of forest cover; changes in
habitat; increase in sediment loads,
streamflow
19th Century
Tanning
Logging
Agriculture (clearing, drainage,
animal agriculture, tillage, crop
rotations, other management
practices)
Canal and dam development
Railroad development
Road development
Rise of urbanization and industrialization
tershed as they have changed over time
from the pre-European settlement period
through modern times. Some, such as canals, road development, urbanization, and
Preferential clearing of hemlock forests,
increased organic loads to watershed
streams
Partial to whole clearing of forests;
corresponding increases in soil
temperature, sediment loads; alterations
of water balance, habitat; modification
of streamflow by damming and logging
operations
Clearing of forest land; increase in
sediment loads, soil nutrient loss,
streamflow, organic nutrient loads
(manure); increase in N fixing crops
(alfalfa)
Introduction of waterborne invasive
species; wetland drainage, alteration of
flow directions of watercourses;
increased nutrient loads to waterways;
changes in water residence time and flow
regime by damming
Increased access to forests and risk of
fire
Increases in impervious surface and
consequent runoff
Increased pollution from sewage and
factory waste
specific agricultural practices, have not been
addressed in detail in this paper, but we
feel it important to acknowledge their influence.
HISTORICAL CHANGES IN POPULATION AND LAND USE
97
Table 1 continued.
20th century
Further dam development for
water supply (in New York
and other municipalities) and
power
Elimination of extreme flows and other
changes in flow regime and sediment
transport; increase in overall water
residence time and evaporation
Development of highways and
other road construction
Increases in impervious surface and
consequent runoff, and related effects;
increased wildlife mortality and other
effects on plant and animal populations
Agricultural decline
Old-field succession, forest regrowth,
expansion of suburban development
Changing agricultural practices
(rise of fertilizer and agrochemicals,
tillage, etc)
Increased inorganic nutrient (fertilizer)
and changes in organic nutrient
(manure) loads
Growth of urban and suburban
development (“sprawl”)
Increased impervious surface, runoff,
pollutants; sediment loads from
construction; stream channelization
Conclusion
Although our analysis can only be considered as incomplete, we have outlined some
of the changes human activities have wrought
on the Hudson River watershed at different
periods of its recent history. Without doubt,
it would require at least one book to discuss all of them in adequate detail. Nevertheless, we hope that our work will remind
readers of the historic legacies shaping the
landscape and suggest to them that current
activities in the watershed will also have
implications for the environment of the region in centuries to come.
Acknowledgments
We thank Christy Goodale, Charlie Canham,
Stephen V. Smith, and two anonymous reviewers for comments on the manuscript
and Stephen V. Smith for advice on GIS
analysis. Partial funding came from National Science Foundation Grant No. DEB0238121 to KL.
References
Aber, J. D., and C. T. Driscoll. 1997. Effects of land
use, climate variation and N deposition on N cycling and C storage in northern hardwood forests.
Global Biogeochemical Cycles 11:639–648.
Abood, K. A., T. L. Englert, S. G. Metzger, C.V. Beckers,
Jr., T. J. Groninger, and S. Mallavaram. 2006.
Current and evolving physical and chemical conditions in the Hudson River estuary. Pages 39–62
in J. R. Waldman, K. E. Limburg, and D. L. Strayer,
editors. Hudson River fishes and their environment. American Fisheries Society, Symposium 51,
Bethesda, Maryland.
Alexander, R. B., and R. A. Smith. 1990. County-level
estimates of nitrogen and phosphorus fertilizer use
in the United States, 1945 to 1985. USGS OpenFile Report 90–130.
Anonymous. 1647. Journal of New Netherland. Trans-
98
SWANEY ET AL.
lated and published in O’Callaghan, E. B. 1851.
The documentary history of the state of New York,
volume IV. Charles Van Benthuysen, Public
Printer, Albany, New York.
Ayres, R. U., and S. R. Rod. 1986. Patterns of pollution in the Hudson-Raritan Basin. Environment
28:14–43.
Battaglin, W. A., and D. A. Goolsby. 1994. U.S. Geological Survey Water-Resources Investigations Report 94–4176.
Bilby, R. E., B. R. Fransen, and P. A. Bisson. 1996.
Incorporation of nitrogen and carbon from spawning coho salmon into the trophic system of small
streams: evidence from stable isotopes. Canadian
Journal of Fisheries and Aquatic Sciences 53:164–
173.
Boyle, R. H. 1969. The Hudson River: a natural and
unnatural history. W. W. Norton and Co., Inc.,
New York.
Brainerd, E., and M. Siegler. 2002. The economic effects of the 1918 influenza epidemic. Centre for
Economic Policy Research, CEPR discussion paper 3791. London. Available: http://
ideas.repec.org/p/cpr/ceprdp/3791.html (December 2004).
Canham, C. D., and S. W. Pacala. 1995. Linking tree
population dynamics and forest ecosystem processes. Pages 84–93 in C. G. Jones and J. H.
Lawton, editors. Linking species and ecosystems.
Chapman and Hall, Inc., New York.
Canham, H. O. 1981. Forest wood products. NYS
Forest Resource Assessment Report No. 7. New
York State Department of Environmental Conservation, Division of Lands and Forests, Forest Resources Planning Program, Albany.
Chapin, F. S. III, B. S. Walker, R. J. Hobbs, D. U.
Hooper, J. H. Lawton, O. E. Sala, and D. Tilman.
1997. Biotic control over the functioning of ecosystems. Science 277:500–504.
Cronon, W. 1983. Changes in the land: Indians, colonists, and the ecology of New England. Hill and
Wang, New York.
Day, G. M. 1953. The Indian as an ecological factor in
the Northeastern forest. Ecology 34:329–346.
Dean, W. E., and E. Gorham. 1998. Magnitude and
significance of carbon burial in lakes, reservoirs,
and peatlands. Geology 26:535–538.
Donaldson, A. L. 1921. A history of the Adirondacks.
The Century Co., New York.
Durbin, A. G., S. W. Nixon, and C. A. Oviatt. 1979.
Effects of the spawning migration of the alewife,
Alosa pseudoharengus, on freshwater ecosystems.
Ecology 60:8–17.
Earle, C., C. Cao, J. Heppen, and S. Otterstrom. 1999.
The Historical United States County Boundary
Files 1790–1999 on CD-ROM. Geoscience Publications. Louisiana State University, Baton Rouge.
Cornell University Press, Ithaca, New York.
Eichel, O. R. 1923. A special report on the mortality
from influenza in New York State during the epidemic of 1918–19. New York State Department
of Health, Albany.
Ellis, D. M., J. A. Frost, H. C. Syrett and H. J. Carman.
1967. A history of New York State. Cornell University Press, Ithaca, New York.
Environmetal Systems Research Institute. 1999.
ARCVIEW GIS 3.2. Environmental Systems Research Institute, Redlands, California.
Evers, A. 1972. The Catskills: from wilderness to
Woodstock. Doubleday & Co., Inc., Garden City,
New York.
Forman, R. T. T., and L. E. Alexander. 1998. Roads
and their major ecological effects. Annual Review
of Ecology and Systematics 29:207–231.
Foster, D. R. 1999. Thoreau’s country. Journey through
a transformed landscape. Harvard University Press,
Cambridge, Massachusettes.
Fox, W. F. 1900. History of the lumber industry in the
state of New York. Pages 237–283 in Sixth Annual Report of the Forest, Fish and Game Commission. James B. Lyon, State Printer, Albany, New
York.
Freeman, M. C., C. M. Pringle, E. A. Greathouse, and
B. J. Freeman. 2003. Ecosystem-level consequences
of migratory faunal depletion caused by dams.
Pages 255–266 in K. E. Limburg and J. R.
Waldman, editors. Biodiversity, status, and conservation of the world’s shads. American Fisheries
Society, Symposium 35, Bethesda, Maryland.
Garman, G. C., and S. A. Macko. 1998. Contribution
of marine-derived organic matter to an Atlantic
coast, freshwater, tidal stream by anadromous clupeid fishes. Journal of the North American
Benthological Society 17:277–285.
Goodale, C. L., and J. D. Aber. 2001. The long-term
effects of land-use history on nitrogen cycling in
northern hardwood forests. Ecological Applications
11:253–267.
Goodale, C. L., J. D. Aber, and W. H. McDowell. 2000.
The long-term effects of disturbance on organic
and inorganic nitrogen export in the White Mountains, New Hampshire. Ecosystems 3:433–450.
Graf, W. L. 1999. Dam nation: a geographic census of
American dams and their large-scale hydrologic
HISTORICAL CHANGES IN POPULATION AND LAND USE
impacts. Water Resources Research 35:1305–1311.
Hamilton, L., B. Askew, and A. Odell. 1980. Forest
history. New York State Forest Resources Assessment:
Report No. 1. Prepared for the New York State
Department of Environmental Conservation, Division of Lands and Forests, Forest Resources Planning Program, Albany.
Hedrick, U. P. 1933. A history of agriculture in the state
of New York. Hill and Wang, New York.
Hobbie, S. E. 1992. Effects of plants on nutrient cycling. Trends in Ecology and Evolution 7:336–339.
Hough, F. B. 1878. First report upon forestry for 1877.
U.S. Government Printing Office, Washington D.C.
Humborg, C., V. Ittekkot, A. Cociasu, and B. Bodungen.
1997. Effect of Danube River dam on Black Sea
biogeochemistry and ecosystem structure. Nature
(London) 386:385–388.
Latty, E. F., C. D. Canham, and P. L. Marks. 2004. The
effects of land-use history on soil properties and
nutrient dynamics in northern hardwood forests of
the Adirondack Mountains. Ecosystems 7:193–
207.
Likens, G. E., and F. H. Bormann. 1995. Biogeochemistry of a forested ecosystem. 2nd edition. SpringerVerlag, New York.
Limburg, K. E., K. A. Hattala, and A. W. Kahnle. 2003.
American shad in its native range. Pages 125–140
in K. E. Limburg and J. R. Waldman, editors.
Biodiversity, status, and conservation of the world’s
shads. American Fisheries Society, Symposium 35,
Bethesda, Maryland.
Lovett, G. M., K. C. Weathers, and M. A. Arthur. 2002.
Control of nitrogen loss from forested watersheds
by soil carbon:nitrogen ratio and tree species composition. Ecosystems 5:712–718.
MacAvoy, S. E., S. A. Macko, S. P. McIninch, and G. C.
Garman. 2000. Marine nutrient contributions to
freshwater apex predators. Oecologia 122:568–
573.
McCully, P. 1996. Silenced rivers: the ecology and politics of large dams. Zed Books, London.
McIsaac, G. F., and X. Hu. 2004. Net N input and
riverine N export from Illinois agricultural watersheds with and without extensive tile drainage. Biogeochemistry 70:253–273.
McMartin, B. 1992. Hides, hemlocks and Adirondack
history. North Country Books, Utica, New York.
Montanus, A. 1671. Description of New Netherland.
Translated from “De Nieuwe en Onbekende
Weereld: of Beschryving van America en ‘t Zuidland”
and published in O’Callaghan, E. B. 1851. The
documentary history of the state of New York, vol-
99
ume IV. Charles Van Benthuysen, Public Printer,
Albany, New York.
Moon, F. F. 1909. Report on the highlands of the Hudson
Forest reservation. Pages 217–233 in Annual reports of the Forest, Fish and Game Commissioner
for 1907–1908-1909. J. B. Lyon, State Printer,
Albany, New York.
Morita, K., and T. Suzuki. 1999. Shifts of food habit
and jaw position of white-spotted charr after damming. Journal of Fish Biology 55:1156–1162.
Nehlsen, W., J. E. Williams, and J. A. Lichatowich. 1991.
Pacific salmon at the crossroads: stocks at risk from
California, Oregon, Idaho, and Washington. Fisheries 16(2):4–21.
New York State Data Center. 2000. Population of New
York State by County 1790 to 1990. New York
State Department of Economic Development, Albany.
New York State Department of Transportation. 2002.
2001 Highway mileage summary. Available: http:/
/ w w w. d o t . s t a t e . n y. u s / t e c h _ s e r v / h i g h /
highwaym.html (December 2004).
Newell, F. H. 1900. A discussion of conservative lumbering and the water-supply. Pages 163–167 in
Sixth Annual Report of the Forest, Fish and Game
Commission. James B. Lyon, State Printer, Albany.
NYCDOH (New York City Department of Health).
1918. Annual Report of the New York City Dept
of Health, New York.
O’Callaghan, E. B., editor. 1850. The documentary history of the state of New York. Volume III. Weed,
Parsons & Co., Albany.
Ollinger, S. V., J. D. Aber, P. B. Reich, and R. J. Freuder.
2002. Interactive effects of nitrogen deposition, tropospheric ozone, elevated CO2 and land use history
on the carbon dynamics of northern hardwood forests. Global Change Biology 8:545–562.
Pacala, S. W., and D. H. Deutschman. 1995. Details
that matter: the spatial distribution of individual
trees maintains forest ecosystem function. Oikos
74:357–365.
Pace, M. L., J. J. Cole, S. R. Carpenter, J. F. Kitchell, J. R.
Hodgson, M. C. Van de Bogert, D. L. Bade, E. S.
Kritzberg, and D. Bastviken. 2004. Whole-lake carbon-13 additions reveal terrestrial support of aquatic
food webs. Nature (London) 427:240–243.
Pringle, C. M. 1997. Exploring how disturbance is transmitted upstream: going against the flow. Journal of
the North American Benthological Society
16:425–438.
Recknagel, A. B. 1923. The forests of New York State.
The Macmillan Company, New York.
100
SWANEY ET AL.
Reed, N. 1875. Early History of Amenia. 1st ed, DeLacey
and Wiley, Amenia, New York. 2nd ed (1964),
Harlem Valley Times, Amenia, New York. 108 p.
Richey, J. E., M. A. Perkins, and C. R. Goldman. 1975.
Effects of kokanee salmon (Oncorhynchus nerka)
decomposition on the ecology of a subalpine stream.
Journal of the Fisheries Research Board of Canada
32:817–820.
Rod, S. R. 1986. Historical reconstruction of pollutant
loadings in river basins. Pages165–170 in R.
Hanham, W. G. Vogt, and M. H. Mickle, editors.
Modeling and simulation, v 17 part 1. Proceedings
of the 17th Annual Pittsburgh Conference.
Rod, S. R., R. U. Ayres, and M. Small. 1989. Reconstruction of historical loadings of heavy metals and
chlorinated hydrocarbon pesticides in the HudsonRaritan basin, 1880–1980. Final Report to the
Hudson River Foundation.
Shorto, R. 2004. The island at the center of the world:
the epic story of Dutch Manhattan and the forgotten colony that shaped America. Doubleday, New
York.
Smith, S. V., W. H. Renwick, J. D. Bartley, and R. W.
Buddemeier. 2002. Distribution and significance
of small, artificial water bodies across the United
States landscape. The Science of the Total Environment 299:21–36.
Stevenson, C. H. 1899. The shad fisheries of the Atlantic coast of the United States. Pages 101–269 in
U.S. Commission of Fish and Fisheries, Part XXIV.
Report of the Commissioner for the year ending
June 30, 1898. Government Printing Office, Washington, D.C.
Stout, L.J. 1958. Atlas of Forestry in New York. State
Univ. College of Forestry at Syracuse University
Swaney. D. P., D. Sherman, and R. W. Howarth. 1996.
Modeling water, sediment and organic carbon discharges in the Hudson-Mohawk Basin: coupling to
terrestrial sources. Estuaries 19:633–647.
U. S. Census Bureau. 2000 Census Data. Available at:
http://www.census.gov (December 2004).
U. S. Census of Agriculture: 1925. United States Department of Commerce, United States Government
Printing Office, Washington DC.
U. S. Census of Agriculture: 1935. United States Department of Commerce, United States Government
Printing Office, Washington DC
U. S. Census of Agriculture: 1945. 1946. volume 1
Counties and State Economic Areas. Part 2. Middle
Atlantic States. United States Department of Commerce, United States Government Printing Office,
Washington DC.
U. S. Census of Agriculture:1954. 1956. volume 1
Counties and State Economic Areas. Part 2. Middle
Atlantic States. United States Department of Commerce, United States Government Printing Office,
Washington DC.
U. S. Census of Agriculture:1964. 1966. volume 1 Statistics for the States and Counties. Part 7. New York.
United States Department of Commerce, United
States Government Printing Office, Washington
DC.
U. S. Census of Agriculture:1974. 1977. volume 1 State
Reports. Part 32. New York. United States Department of Commerce, United States Government
Printing Office, Washington DC.
U. S. Census of Agriculture:1987. 1989. volume 1
Geographic Area Series. Part 32. New York State
and County Data. United States Department of
Commerce, United States Government Printing
Office, Washington DC.
United States Department of Agriculture, Forest Service. 1970–1977. Forest Resource Bulletins NE20 (NY-1968), NE-34 (NJ-1972), NE-44 (CT1972), NE-46 (VT-1973), NE-48 (MA-1972).
United States Department of Agriculture, Forest Service. 1990–1999. Forest Resource Inventory Reports NE-112 (NJ-1987), NE-132 (NY-1993),
NE-145 (VT-1997), NE-147(CT-1998), NE148 (MA-1998).
USEPA. 2001. Better Assessment Science Integrating
Point and Nonpoint Sources. BASINS Version 3.0.
Users’ Manual. Office of Water. United States Environmental Protection Agency Report EPA
823B01001. Washington.
Vörösmarty, C. J., K. P. Sharma, B. M. Fekete, A. H.
Copeland, J. Holden, J. Marble, and J. A. Lough.
1997. The storage and aging of continental runoff
in large reservoir systems of the world. Ambio
26:210–219.
Waring, R. H., and S. W. Running. 1998. Forest Ecosystems: Analysis at Multiple Scales. 2nd edition. Academic Press, San Diego.
Weaver, M. 1964. History of Tile Drainage (in America
prior to 1900). M. M. Weaver, Waterloo, NY.
Whipple, J. S. 1909. Report of the Superintendent of
Forests for the year 1909. State of New York Forest,
Fish and Game Commission, Annual Reports
1907–1908-1909. J. B. Lyon Company, Albany.
Yanai, R. D., M. A. Arthur, T. G. Siccama, and C. A.
Federer. 2000. Challenges of measuring forest floor
organic matter dynamics: repeated measures from a
chronosequence. Forest Ecology and Management
138:273–283.
State
CT
MA
NJ
NJ
NJ
NJ
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
County
Fairfield
Berkshire
Bergen
Hudson
Passaic
Sussex
Albany
Bronx
Columbia
Delaware
Dutchess
Essex
Fulton
Greene
Hamilton
Herkimer
Lewis
Madison
Montgomery
New York
1621
2411
606
122
479
1349
1357
109
1647
3745
2077
4654
1285
1678
4457
3657
3305
1699
1049
73
Area
(km2)
Upper Hudson
0.214
0.414
Sacandaga
0.273
0.291
0.010
0.363
0.145
0.462
0.015
0.030
0.834 0.165
0.596
0.956
0.000
0.103
0.709
0.796
Hudson-Hoosic
0.063 0.141
Mohawk
0.000
Schoharie
0.040
Middle Hudson
0.158
Rondout
0.036
0.360
Hudson-Wappinger
0.662
0.000
Lower Hudson
0.988
0.044
0.175
0.041
0.018
0.064
0.064
0.198
0.041
0.018
0.036
0.360
1.000
0.175
0.956
0.010
0.809
0.414
1.000
0.959
0.632
0.462
0.015
0.030
1.000
0.988
total
Table A.1 Modern (1990) values of county areas and the proportions falling within the Hudson-Mohawk watershed and its hydrological catalogue units (HUCs).
HISTORICAL CHANGES IN POPULATION AND LAND USE
101
State
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
VT
VT
VT
County
Oneida
Orange
Otsego
Putnam
Rensselaer
Rockland
Saratoga
Schenectady
Schoharie
Sullivan
Ulster
Warren
Washington
Westchester
Bennington
Rutland
Windham
Table A.1 continued
3142
2113
2598
601
1694
451
2103
534
1611
2512
2919
2253
2165
1121
1751
2414
2044
Area
(km2)
Hudson-Hoosic
0.708
0.002
0.001
0.189 0.055
0.512
0.543
Upper Hudson
0.272 0.580
Sacandaga
0.014
0.406
Mohawk
Schoharie
0.134
0.619 0.130
0.040 0.805
0.000
0.010 0.006
0.468
Middle Hudson
0.317
0.252
0.067
0.594
Rondout
0.090
0.459
0.456
0.988
Hudson-Wappinger
0.084
0.001
0.181
0.312
0.988
Lower Hudson
0.662
0.365
0.818
0.018
total
0.468
0.474
0.015
1.000
1.000
0.366
1.000
1.000
0.912
0.090
0.860
0.787
0.512
0.662
0.708
0.002
0.001
102
SWANEY ET AL.
1
CT
MA
NJ
NJ
NJ
NJ
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
Fairfield
Berkshire
Bergen
Hudson
Passaic
Sussex
Albany
Bronx
Columbia
Delaware
Dutchess
Essex1
Fulton
Greene
Hamilton
Herkimer
Kings
Lewis
Madison
Montgomery
New York
Oneida
Orange
Otsego
Putnam
Queens
Rensselaer
Rockland
0.065
0.197
0.038
0.000
0.000
0.219
0.999
0.000
0.956
0.000
0.852
0.201
0.000
0.000
0.000
0.000
0.000
0.000
0.000
0.167
0.470
0.000
0.648
0.000
0.000
0.000
0.000
0.000
1790
0.058
0.179
0.036
0.015
0.030
0.324
0.903
0.864
0.009
0.731
0.382
0.903
0.866
0.567
0.418
0.000
0.013
0.026
0.903
0.358
0.423
0.710
0.013
0.903
0.000
0.903
0.333
0.864
0.009
0.732
0.381
0.000
0.867
0.567
0.418
0.000
0.013
0.026
0.903
0.358
0.423
0.710
0.013
0.903
0.000
0.903
0.333
1850
0.058
0.178
0.034
0.000
0.000
0.213
0.903
1820
1790 values for Clinton County, from which Essex County was later formed.
State
County
0.864
0.009
0.731
0.380
0.903
0.866
0.566
0.418
0.000
0.013
0.027
0.903
0.367
0.422
0.710
0.014
0.903
0.000
0.903
0.331
0.058
0.179
0.037
0.016
0.032
0.326
0.903
1880
Year
0.058
0.179
0.037
0.016
0.032
0.326
0.903
0.161
0.864
0.009
0.731
0.380
0.903
0.866
0.566
0.418
0.000
0.013
0.027
0.903
0.892
0.422
0.428
0.014
0.903
0.000
0.903
0.331
1920
0.064
0.198
0.041
0.018
0.036
0.360
1.000
0.175
0.956
0.010
0.809
0.414
1.000
0.959
0.632
0.462
0.000
0.015
0.030
1.000
0.988
0.468
0.474
0.015
1.000
0.000
1.000
0.366
1950
0.064
0.198
0.041
0.018
0.036
0.360
1.000
0.175
0.956
0.010
0.809
0.414
1.000
0.959
0.632
0.462
0.000
0.015
0.030
1.000
0.988
0.468
0.474
0.015
1.000
0.000
1.000
0.366
1970
0.064
0.198
0.041
0.018
0.036
0.360
1.000
0.175
0.956
0.010
0.809
0.414
1.000
0.959
0.632
0.462
0.000
0.015
0.030
1.000
0.988
0.468
0.474
0.015
1.000
0.000
1.000
0.366
2000
Table A.2 Estimated proportions of county areas falling within the Hudson-Mohawk watershed over time (to 3 decimal places).
HISTORICAL CHANGES IN POPULATION AND LAND USE
103
State
Saratoga
NY
Schenectady NY
Schoharie
NY
Sullivan
NY
Ulster
NY
Warren
NY
Washington NY
Westchester NY
Bennington VT
Rutland
VT
Windham
VT
County
Table A.2 continued
0.000
0.000
0.000
0.000
0.454
0.000
0.617
0.661
0.707
0.002
0.001
1790
0.903
0.903
0.821
0.082
0.777
0.710
0.464
0.597
0.641
0.003
0.001
1820
0.903
0.903
0.823
0.082
0.777
0.710
0.464
0.597
0.639
0.001
0.001
1850
0.903
0.904
0.824
0.082
0.777
0.711
0.463
0.598
0.639
0.001
0.001
1880
Year
0.903
0.904
0.824
0.082
0.777
0.711
0.463
0.598
0.639
0.001
0.001
1920
1.000
1.000
0.912
0.090
0.860
0.787
0.512
0.662
0.708
0.002
0.001
1950
1.000
1.000
0.912
0.090
0.860
0.787
0.512
0.662
0.708
0.002
0.001
1970
1.000
1.000
0.912
0.090
0.860
0.787
0.512
0.662
0.708
0.002
0.001
2000
104
SWANEY ET AL.
1
CT
MA
NJ
NJ
NJ
NJ
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
Fairfield
Berkshire
Bergen
Hudson
Passaic
Sussex
Albany
Bronx
Columbia
Delaware
Dutchess
Essex2
Fulton
Greene
Hamilton
Herkimer
Kings
Lewis
Madison
Montgomery
New York
36290
30263
12601
19500
75980
27496
45276
16151
4549
28852
33111
1790
38330
26587
46615
12811
32752
38116
22996
1251
31017
11187
9227
32208
37569
123706
-
-
-
42739
35720
18178
1820
1790 values for Clinton County, from which Essex County was later formed.
State
County
Table A.3 Population of counties of Hudson/Mohawk watershed.
43073
39834
58992
31148
20171
33126
2188
38244
138882
24564
43072
31992
515547
-
59775
49591
14725
21822
22569
22989
93279
1850
112042
69032
36786
187944
68860
23539
154890
47928
42721
79184
34515
30985
32695
3923
42669
599495
31416
44112
38315
1206299
1880
Year
1950
320936
504342
113033
132966
210703
539139
629154
647437
259174
337093
24905
34423
186106
239386
732016 1451277
38930
43182
42774
44420
91747
136781
31871
35086
44927
51021
25796
28745
3970
4105
64962
61407
2018356 2738175
23704
22521
39535
46214
57928
59594
2284103
1960101
1920
792814
149402
897148
607839
460782
77528
286742
1471701
51519
44718
222295
34631
52637
33136
4714
67633
2602012
23644
62864
55883
1539233
1970
882567
134953
884118
608975
489049
144166
294565
1332650
63094
48055
280150
38851
55073
48195
5379
64427
2465326
26944
69441
49708
1537195
2000
HISTORICAL CHANGES IN POPULATION AND LAND USE
105
State
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
VT
VT
VT
County
Oneida
Orange
Otsego
Putnam
Queens
Rensselaer
Rockland
Saratoga
Schenectady
Schoharie
Sullivan
Ulster
Warren
Washington
Westchester
Bennington
Rutland
Windham
Table A.3 continuted
18477
16013
29370
14077
23978
12206
15590
17572
1790
50997
41213
44856
11268
21519
40153
8837
36052
13081
23154
8900
30934
9453
38831
32638
16125
29983
28457
1820
99566
57145
48638
14138
36833
73363
16962
45646
20054
33548
25088
59384
17199
44750
58263
49591
33059
29062
1850
115475
88220
51397
15181
90574
115328
27690
55156
23538
32910
32491
85838
25179
47871
108988
21950
41829
26763
1880
Year
182833
119844
46200
10802
469042
113129
45548
60029
109363
21303
33163
74979
31673
44888
344436
21577
46213
26373
1920
222855
152255
50763
20307
1550849
132607
89276
74869
142497
22703
40731
92621
39205
47144
625816
24115
45905
28749
1950
273037
221657
56181
56696
1987174
152510
229903
121764
161078
24750
52580
141241
49402
52725
894406
29282
52637
33476
1970
235469
341367
61676
95745
2229379
152538
286753
200635
146555
31582
73966
177749
63303
61042
923459
36994
63400
44216
2000
106
SWANEY ET AL.
1820
10197
16428
67504
76661
33810
113066
37214
43664
80492
479035
1790
Upper Hudson
5150
Sacandaga
7783
Hudson-Hoosic
37912
Mohawk
9441
Schoharie
15490
Middle Hudson
63095
Rondout
16710
Hudson-Wappinger 30956
Lower Hudson
46202
Whole Watershed 232740
Subwatershed
21281
19965
111102
120588
54873
192406
58515
58177
243235
880142
1850
26891
29140
26974
32952
118562
125303
147749
258877
63452
69693
274192
302450
84538
99732
81241
96137
543088 2426964
1366688 3441248
1880
Time
1920
Table A.4 Estimated population of watersheds (HUCs) of Hudson/Mohawk watershed.
37708
43732
161731
335476
91451
414075
140184
149582
2727641
4101579
1950
43833
59062
206697
381366
103184
503923
215218
238645
2611766
4363694
1970
54356
84051
260679
361641
112403
543985
313554
324474
2667309
4722454
2000
HISTORICAL CHANGES IN POPULATION AND LAND USE
107
Fairfield
Berkshire
Bergen
Hudson
Passaic
Sussex
Albany
Bronx
Columbia
Delaware
Dutchess
Essex
Fulton
Greene
Hamilton
Herkimer
Kings
Lewis
Madison
Montgomery
New York
County
CT
MA
NJ
NJ
NJ
NJ
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
State
1072
1811
440
69
390
989
1204
0
1455
2610
1923
1229
666
1272
152
1377
84
943
1574
968
11
1850
993
1561
426
11
387
937
1265
0
1510
3073
1862
1848
1066
1372
397
1540
49
1757
1576
980
5
1870
1266
1944
380
12
352
1027
1239
0
1653
2982
1986
1692
1148
1439
402
2173
42
1874
1595
1160
10
1880
Year
913
1442
150
3
130
868
1126
5
1378
3015
1767
1257
794
1144
147
1353
4
1769
1467
919
1
1920
Table A.5 Land in farms (km2) in counties of Hudson/Mohawk watershed.
391
818
54
4
38
663
651
1
1052
2550
1229
796
373
695
30
1142
1
1298
1285
828
0
1950
92
430
26
0
6
398
349
0
706
1472
656
295
197
284
0
801
0
887
955
653
0
1969
48
254
11
0
9
295
230
0
465
743
432
195
139
197
3
574
0
727
752
546
0
1997
108
SWANEY ET AL.
Oneida
Orange
Otsego
Putnam
Queens
Rensselaer
Rockland
Saratoga
Schenectady
Schoharie
Sullivan
Ulster
Warren
Washington
Westchester
Bennington
Rutland
Windham
County
State
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
VT
VT
VT
Table A.5 continued.
2696
1715
2218
489
687
1415
313
1671
428
1259
956
1785
898
1627
1020
906
1800
1697
1850
2645
1692
2454
515
624
1452
227
1661
501
1511
1599
1860
1405
1884
983
771
1786
1757
1870
2871
2023
2594
563
673
1718
321
1867
508
1602
1909
2188
1361
1987
1035
1103
1923
1768
1880
2475
1418
2320
457
57
1245
162
1363
415
1396
1679
1654
867
1757
421
839
1766
1342
1920
1963
1108
1938
174
2
921
70
811
245
1170
777
921
298
1558
196
517
1361
892
1950
Year
1294
636
1318
56
0
493
16
401
140
699
375
371
43
1082
61
251
796
334
1969
401
2
295
74
448
235
279
37
789
30
131
509
189
875
384
838
14
1997
HISTORICAL CHANGES IN POPULATION AND LAND USE
109
Upper Hudson
Sacandaga
Hudson-Hoosic
Mohawk
Schoharie
Middle Hudson
Rondout
Hudson-Wappinger
Lower Hudson
Whole Watershed
Watershed
957
776
3032
3520
1715
4498
1862
1848
1264
19472
1850
1486
1027
3065
3930
1954
4703
1918
1816
1233
21133
1870
1408
1093
3583
4513
2066
5113
2244
2016
1343
23379
1880
Year
1256
888
3288
3751
1922
4764
2053
1837
1113
20872
1920
633
504
2452
3083
1647
3584
1425
1296
260
14507
1950
1969
309
308
1784
2484
1215
2533
1081
1072
238
10667
Table A.6 Estimated land in farms (km2) in subwatersheds (HUCs) of Hudson/Mohawk watershed.
123
143
1011
1696
682
1351
496
510
79
5946
1997
110
SWANEY ET AL.
111
HISTORICAL CHANGES IN POPULATION AND LAND USE
Table A.7 Forested lands (km2) in counties of Hudson/Mohawk watershed. – indicates no data. 1968
and 1993 include estimates based on reports from succeeding years
County
Fairfield
Berkshire
Bergen
Hudson
Passaic
Sussex
Albany
Bronx
Columbia
Delaware
Dutchess
Essex
Fulton
Greene
Hamilton
Herkimer
Kings
Lewis
Madison
Montgomery
New York
Oneida
Orange
Otsego
Putnam
Queens
Rensselaer
Rockland
Saratoga
Schenectady
Schoharie
Sullivan
Ulster
Warren
Washington
Westchester
Bennington
Rutland
Windham
State
CT
MA
NJ
NJ
NJ
NJ
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
NY
VT
VT
VT
Year
1875
1920
1955
1968
1993
245
0
251
1572
399
1651
539
581
2233
537
8
1603
366
146
0
823
611
620
210
220
336
241
600
93
396
1348
1409
878
513
261
-
324
0
324
1578
405
3845
708
688
4249
2509
0
1862
304
101
0
951
1214
648
283
0
445
324
911
142
405
1518
1740
1680
708
142
-
384
0
638
1974
955
4115
888
1000
4422
2410
0
2261
464
194
0
1151
1123
905
418
0
746
333
1172
122
732
1697
2334
2074
884
598
-
756
789
539
0
842
2492
1071
4322
994
1182
4374
2776
0
2518
776
264
0
1581
1026
1355
398
0
996
0
1350
190
871
1813
2167
2120
1062
406
1522
1837
2022
600
850
691
0
1016
2708
1121
4184
987
1329
4367
2776
0
2497
806
394
0
1843
1159
1493
417
0
1034
210
1423
283
1081
1968
2359
2159
1232
487
1529
1990
1765
SWANEY ET AL.
112
Table A.8 Estimated forested land (km2) in subwatersheds of Hudson/Mohawk watershed.
Watershed
Upper Hudson
Sacandaga
Hudson-Hoosic
Mohawk
Schoharie
Middle Hudson
Rondout
Hudson-Wappinger
Lower Hudson
Whole Watershed
Year
1875
1920
1955
1477
996
718
1513
559
1373
947
552
417
8551
3078
1756
1051
2805
610
1716
1347
763
436
13562
3790
2175
1549
3428
1078
2874
1738
1254
921
18807
1960s-70s
1980s-90s
3892
2249
2931
4024
1305
3436
1912
1283
708
21740
3856
2273
3082
4337
1585
3937
2096
1377
848
23390
Table A.9 Numbers of dams in each hydrologic subunit of the Hudson River basin.
1
Hydrologic
Area
Upper Hudson
Sacandaga
Hudson-Hoosic
Mohawk
Schoharie
Middle Hudson
Rondout
Hudson-Wappingers
Lower Hudson
Bronx2
4222
2720
4869
6605
2401
6190
3082
2403
1864
(?)
Number of dams
1
0
0
0
0
0
0
1
0
0
2
2
3
6
4
2
3
4
5
0
34
15
35
37
22
51
61
45
62
0
16
8
25
31
17
50
41
40
37
1
Includes only those dams falling within the Hudson part of this hydrologic subunit.
2
3
12
16
1
13
6
11
14
1
0
3
7
17
6
7
2
3
11
1
55
31
82
107
50
123
113
104
129
3
Download